Elsevier

Vitamins & Hormones

Volume 79, 2008, Pages 267-292
Vitamins & Hormones

Chapter 9 Regulation of Human Dihydrofolate Reductase Activity and Expression

https://doi.org/10.1016/S0083-6729(08)00409-3Get rights and content

Abstract

Dihydrofolate reductase (DHFR) enzyme catalyzes tetrahydrofolate regeneration by reduction of dihydrofolate using NADPH as a cofactor. Tetrahydrofolate and its one carbon adducts are required for de novo synthesis of purines and thymidylate, as well as glycine, methionine and serine. DHFR inhibition causes disruption of purine and thymidylate biosynthesis and DNA replication, leading to cell death. Therefore, DHFR has been an attractive target for chemotherapy of many diseases including cancer. Over the following years, in order to develop better antifolates, a detailed understanding of DHFR at every level has been undertaken such as structure‐functional analysis, mechanisms of action, transcriptional and translation regulation of DHFR using a wide range of technologies. Because of this wealth of information created, DHFR has been used extensively as a model system for enzyme catalysis, investigating the relations between structure in‐silico structure‐based drug design, transcription from TATA‐less promoters, regulation of transcription through the cell cycle, and translational autoregulation. In this review, the current understanding of human DHFR in terms of structure, function and regulation is summarized.

Introduction

The enzyme dihydrofolate reductase (DHFR, 5,6,7,8‐tetrahydrofolate:NADP+ oxidoreductase, EC 1.51.3) catalyzes the reduction of dihydrofolate (H2F) to tetrahydrofolate (H4F) utilizing NADPH as a cofactor. H4F and its derivatives are essential cofactors in the synthesis of thymidylate, purines, and some amino acids (Figure 9.1, Figure 9.2) (Blakley and Cocco, 1984, Futterman, 1957, Osborn and Huennekens, 1958). Inhibition of DHFR results in a depletion of the reduced folate pools, inhibition of DNA synthesis, and cell death. Due to its biological significance, DHFR has proven to be an important target of antineoplastic, antiprotozoal, antifungal, and antimicrobial drugs in addition to its use for the treatment of other nonmalignant diseases, such as arthritis.

Antifolates are the oldest of the antimetabolite class of anticancer drugs and have been used in the clinic for more than four decades. The first clinically useful antifolate was aminopterin, a tight binding inhibitor of DHFR. Treatment with aminopterin led to the first‐ever remissions in childhood leukemia. Soon after, methotrexate (MTX) replaced aminopterin based on animal studies showing that MTX had a better therapeutic index.

Over the following years, in order to develop better antifolates, a detailed understanding of DHFR at every level has been undertaken such as structure–functional analysis, mechanisms of action, transcriptional and translation regulation of DHFR using a wide range of technologies. Because of this wealth of information created, DHFR has been used extensively as a model system for enzyme catalysis, investigating the relations between structure in silico structure‐based drug design, transcription from TATA‐less promoters, regulation of transcription through the cell cycle, and translational autoregulation.

In this review, the current understanding of human DHFR is summarized. We begin with the structure and kinetic mechanism of enzyme of DHFR. The review then concentrates on the genomic organization, polymorphisms, transcriptional, and translational regulation of DHFR. We refer readers to earlier reviews that discuss many aspects of DHFR regulation for additional in‐depth analysis (Azizkhan et al., 1993, Banerjee et al., 2002, Blakley, 1995, Cody et al., 2006, Hammes‐Schiffer and Benkovic, 2006, Schnell et al., 2004, Slansky and Farnham, 1996).

Section snippets

Structure and Binding of Dihydrofolate, MTX, and NADPH

As a result of its importance, a detailed picture how DHFR works emerged by studying its kinetics and structure; however, the most extensive structural characterization has been done for Escherichia coli. Fortunately, the amino acids required for catalysis and the general features of the secondary structures have been conserved throughout the evolution expect in protozoa and plants where DHFR is fused to thymidylate synthase as a bifunctional enzyme.

DHFR shares a modified version of a common

Mechanism of DHFR Catalysis

DHFR catalyzes the reduction of 7,8‐dihydrofolate to 5,6,7,8‐tetrahydrofolate using NADPH as the hydride donor. Specifically, the pro‐R hydrogen of NADPH is transferred to C6 of the pteridine ring with concomitant protonation at the N5 position (Fig. 9.2A) (Benkovic and Hammes‐Schiffer, 2003). The kinetic mechanism of E. coli and human DHFR catalysis are studied extensively by several groups (Blakley, 1995, Hammes‐Schiffer and Benkovic, 2006, Sawaya and Kraut, 1997, Schnell et al., 2004).

Alternative Substrates: Folic Acid and Dihydrobiopterin

H2F and its polyglutamylate forms are the major substrate of DHFR, although the fully oxidized folate and biopterin are poor substrates of DHFR. In January 1998, Food and Drug Administration initiated the folic acid fortification program, which requires addition of 0.43–1.4 mg folic acid per pound to enriched flour. For folic acid to be converted to the physiologically useful form found in the blood stream, 5‐methyltetrahydrofolate, folic acid is reduced to H4F in the upper small intestine or

Genomic Organization of DHFR

The functional DHFR gene (GeneID: 1719) is located at chromosome 5q11.2‐q13.3. It was cloned, mapped, and sequenced by the Nienhuis and Attardi laboratories (Chen et al., 1982, Chen et al., 1984, Masters and Attardi, 1983, Maurer et al., 1984, Maurer et al., 1985). The DHFR gene is ∼30 kb long and contains six exons and five introns with strictly conserved intron/exon boundaries. The approximate length of the introns varies between 0.35 kb (intron 1) and 11.4 kb (intron 3). After pre‐mRNA

Human Dihydrofolate Reductase Pseudogenes

There is one functional DHFR gene located in the region of q11.1‐q13.3 region of chromosome 5 and at least four pseudogenes interspersed on several chromosomes. Three of the DHFR pseudogenes are located on chromosomes 3, 6, and 18 (Anagnou et al., 1984, Blakley and Sorrentino, 1998, Maurer et al., 1984, Maurer et al., 1985, Polymeropoulos et al., 1991). These pseudogenes presumably have no activity, since unlike hDHFR they were not amplified in MTX‐resistant cell lines (Anagnou et al., 1984,

Transcriptional Regulation

DHFR expression is regulated at many levels. Cellular transcription is regulated not only by the transcription factors, but also through chromatin remodeling involving acetylation and methylation of histones. Histone acetyl transferase (HAT) may be recruited to the DNA after transcriptional factor binding, thereby enhancing nucleosomal relaxation followed by increased transcription. Furthermore, transcription factors can bind directly to histone deacetylases (HDAC) that remove acetyl groups

Polymorphisms of DHFR

Single‐nucleotide polymorphisms (SNPs) are present in 1% of the human genome. SNPs might occur in coding region or noncoding region; the former may lead to defective protein due to amino acid changes, the latter may affect the gene transcription, RNA splicing, or RNA stability.

To date, no polymorphism within the coding region of DHFR have been found possibly due to the critical role of enzyme (Banerjee et al., 2002, Blakley and Sorrentino, 1998, Gellekink et al., 2007, Parle‐McDermott et al.,

Posttranscriptional Regulation of DHFR

Recently, a new posttranscriptional mechanism of gene regulation has been found in mammalian cells and plants. microRNAs (miRNA) are 22 nucleotide noncoding RNAs and act by translational repression, mRNA cleavage, mRNA deadenylation, or transcriptional silencing. miRNAs share the RNAi machinery to form hybrids with target mRNA by anchoring the 3′ end of RNA (Murchison and Hannon, 2004). The 3′UTR of DHFR harbors a mir‐24 microRNA‐binding site which is next to the SNP 829C > T (vide supra). The

Translational Regulation of DHFR

It has been over four decades since initial reports described a rapid increase in DHFR levels in response to the antifolate, MTX (Bertino et al., 1962). As increased intracellular levels of DHFR may hinder clinical success, considerable efforts and attention have been directed toward determining the exact molecular mechanism of this induction. As a result of these efforts, we and others have determined that a translational mechanism underlies this rapid increase. Recently, it has been shown

Acknowledgments

We regret that many important references could not be cited, or were cited indirectly by citing review particles due to space limitations. This work was supported by Grant CA 08010 from the United States Public Health Service (to J. R. B.) and Department of Medicine Grant from UMDNJ (to E. A. E.). We gratefully acknowledge helpful discussions with Dr. Joseph Bertino, Dr. Debabrata Banerjee, and Dr. Vivian Cody for providing the ribbon diagram of human dihydrofolate reductase.

References (102)

  • W.G. Johnson et al.

    Common dihydrofolate reductase 19‐base pair deletion allele: A novel risk factor for preterm delivery

    Am. J. Clin. Nutr.

    (2005)
  • J.N. Masters et al.

    The nucleotide sequence of the cDNA coding for the human dihydrofolic acid reductase

    Gene

    (1983)
  • J.N. Masters et al.

    A human dihydrofolate reductase pseudogene and its relationship to the multiple forms of specific messenger RNA

    J. Mol. Biol.

    (1983)
  • E.P. Murchison et al.

    miRNAs on the move: miRNA biogenesis and the RNAi machinery

    Curr. Opin. Cell Biol.

    (2004)
  • V. Noe et al.

    An intron is required for dihydrofolate reductase protein stability

    J. Biol. Chem.

    (2003)
  • M.J. Osborn et al.

    Enzymatic reduction of dihydrofolic acid

    J. Biol. Chem.

    (1958)
  • K.K. Park et al.

    Modulation of Sp1‐dependent transcription by a cis‐acting E2F element in dhfr promoter

    Biochem. Biophys. Res. Commun.

    (2003)
  • T. Shimada et al.

    A human dihydrofolate reductase intronless pseudogene with an Alu repetitive sequence: Multiple DNA insertions at a single chromosomal site

    Gene

    (1984)
  • T. Shimada et al.

    Chromatin structure of the human dihydrofolate reductase gene promoter. Multiple protein‐binding sites

    J. Biol. Chem.

    (1986)
  • N. Skacel et al.

    Identification of amino acids required for the functional up‐regulation of human dihydrofolate reductase protein in response to antifolate Treatment

    J. Biol. Chem.

    (2005)
  • I.J. van der Linden et al.

    Variation and expression of dihydrofolate reductase (DHFR) in relation to spina bifida

    Mol. Genet. Metab.

    (2007)
  • A. Watanabe et al.

    Genomic organization and expression of the human MSH3 gene

    Genomics

    (1996)
  • H.T. Abelson et al.

    Kinetics of tetrahydrobiopterin synthesis by rabbit brain dihydrofolate reductase

    Biochem. J.

    (1978)
  • N.P. Anagnou et al.

    Chromosomal organization of the human dihydrofolate reductase genes: Dispersion, selective amplification, and a novel form of polymorphism

    Proc. Natl. Acad. Sci. USA

    (1984)
  • N.P. Anagnou et al.

    Chromosomal localization and racial distribution of the polymorphic human dihydrofolate reductase pseudogene (DHFRP1)

    Am. J. Hum. Genet.

    (1988)
  • J.C. Azizkhan et al.

    Nucleotide sequence and nuclease hypersensitivity of the Chinese hamster dihydrofolate reductase gene promoter region

    Biochemistry

    (1986)
  • J.C. Azizkhan et al.

    Transcription from TATA‐less promoters: Dihydrofolate reductase as a model

    Crit. Rev. Eukaryot. Gene Expr.

    (1993)
  • D. Banerjee et al.

    Role of E2F‐1 in chemosensitivity

    Cancer Res.

    (1998)
  • D. Banerjee et al.

    Levels of E2F‐1 expression are higher in lung metastasis of colon cancer as compared with hepatic metastasis and correlate with levels of thymidylate synthase

    Cancer Res.

    (2000)
  • S.J. Benkovic et al.

    A perspective on enzyme catalysis

    Science

    (2003)
  • J.R. Bertino et al.

    Increased level of dihydrofolic reductase in leucocytes of patients treated with amethopterin

    Nature

    (1962)
  • J.R. Bertino et al.

    The “induction” of dihydrofolic reductase activity in leukocytes and erythrocytes of patients treated with amethopterin

    J. Clin. Invest.

    (1963)
  • J.R. Bertino et al.

    “Induction” of dihydrofolate reductase: Purification and properties of the “induced” human erythrocyte and leukocyte enzyme and normal bone marrow enzyme

    Cancer Res.

    (1970)
  • R.L. Blakley

    Eukaryotic dihydrofolate reductase

    Adv. Enzymol. Relat. Areas Mol. Biol.

    (1995)
  • R.L. Blakley et al.

    Dismutation of dihydrofolate by dihydrofolate reductase

    Biochemistry

    (1984)
  • R.L. Blakley et al.

    In vitro mutations in dihydrofolate reductase that confer resistance to methotrexate: Potential for clinical application

    Hum. Mutat.

    (1998)
  • S.W. Blume et al.

    The 5′‐untranslated RNA of the human dhfr minor transcript alters transcription pre‐initiation complex assembly at the major (core) promoter

    J. Cell. Biochem.

    (2003)
  • L.D. Botto et al.

    5,10‐Methylenetetrahydrofolate reductase gene variants and congenital anomalies: A HuGE review

    Am. J. Epidemiol.

    (2000)
  • C.A. Bottoms et al.

    A structurally conserved water molecule in Rossmann dinucleotide‐binding domains

    Protein Sci.

    (2002)
  • C. Bystroff et al.

    Crystal structure of unliganded Escherichia coli dihydrofolate reductase. Ligand‐induced conformational changes and cooperativity in binding

    Biochemistry

    (1991)
  • O. Carugo et al.

    NADP‐dependent enzymes. I: Conserved stereochemistry of cofactor binding

    Proteins

    (1997)
  • O. Carugo et al.

    NADP‐dependent enzymes. II: Evolution of the mono‐ and dinucleotide binding domains

    Proteins

    (1997)
  • K. Chalupsky et al.

    Endothelial dihydrofolate reductase: Critical for nitric oxide bioavailability and role in angiotensin II uncoupling of endothelial nitric oxide synthase

    Proc. Natl. Acad. Sci. USA

    (2005)
  • Y.C. Chang et al.

    Cooperation of E2F‐p130 and Sp1‐pRb complexes in repression of the Chinese hamster dhfr gene

    Mol. Cell. Biol.

    (2001)
  • M.J. Chen et al.

    Intronless human dihydrofolate reductase genes are derived from processed RNA molecules

    Proc. Natl. Acad. Sci. USA

    (1982)
  • E. Chu et al.

    Specific binding of human dihydrofolate reductase protein to dihydrofolate reductase messenger RNA in vitro

    Biochemistry

    (1993)
  • V. Cody et al.

    Structural characteristics of antifolate dihydrofolate reductase enzyme interactions

    Crystallogr. Rev.

    (2006)
  • V. Cody et al.

    Crystal structure determination at 2.3 A of recombinant human dihydrofolate reductase ternary complex with NADPH and methotrexate‐gamma‐tetrazole

    Anticancer Drug Des.

    (1992)
  • V. Cody et al.

    Comparison of ternary crystal complexes of F31 variants of human dihydrofolate reductase with NADPH and a classical antitumor furopyrimidine

    Anticancer Drug Des.

    (1998)
  • K.H. Cowan et al.

    Regulation of dihydrofolate reductase in human breast cancer cells and in mutant hamster cells transfected with a human dihydrofolate reductase minigene

    Mol. Pharmacol.

    (1986)
  • Cited by (45)

    • Pemetrexed-conjugated hyaluronan for the treatment of malignant pleural mesothelioma

      2019, European Journal of Pharmaceutical Sciences
      Citation Excerpt :

      In previous reports, MTX conjugated with, e.g., dendrimer (Gurdag et al., 2006; Li et al., 2012) and PEG (Riebeseel et al., 2002) showed inhibitory activity against DHFR, yet their activities were less than native MTX, which was consistent with the case of HA-ADH-PMX. It has been reported that PMX binds to DHFR with its 2-amino group and N3 of the pteridine ring (Abali et al., 2008; Singh et al., 2018), both of which were preserved in HA-ADH-PMXs. Therefore, the decrease in the inhibitory activity would be explained by e.g., mobility of the binding moieties.

    • Evaluation of drug candidature of some anthraquinones from Morinda citrifolia L. as inhibitor of human dihydrofolate reductase enzyme: Molecular docking and in silico studies

      2017, Computational Toxicology
      Citation Excerpt :

      Prevention research is a promising approach to reduce cancer incidence. In particular, chemoprevention using herbal medicine has been emerging due to its anticancer activity [4]. Morinda citrifolia L. (Rubiaceae), known as ‘‘Noni”, is a shrub which grows in tropical Asia and Polynesia.

    • Targeting nuclear thymidylate biosynthesis

      2017, Molecular Aspects of Medicine
      Citation Excerpt :

      One of the first successful and still widely used antifolates, methotrexate (MTX), targets DHFR (Bertino et al., 1962). Binding of MTX to DHFR inhibits DHFR activity, and also disrupts the interaction between DHFR and its own mRNA, relieving the repressed DHFR translation and leading to increased DHFR protein levels (Abali et al., 2008; Appleman et al., 1988; Bystroff and Kraut, 1991). Many cells develop resistance to MTX due to de-repression of DHFR translation by preventing DHFR from binding to its own mRNA (Bertino et al., 1962).

    View all citing articles on Scopus
    View full text